Next Article in Journal
Identification and Spatiotemporal Expression of a Putative New GABA Receptor Subunit in the Human Body Louse Pediculus humanus humanus
Next Article in Special Issue
The Function and Mechanism of Long Noncoding RNAs in Adipogenic Differentiation
Previous Article in Journal
Deafness DFNB128 Associated with a Recessive Variant of Human MAP3K1 Recapitulates Hearing Loss of Map3k1-Deficient Mice
Previous Article in Special Issue
A Characterization and Functional Analysis of Peroxisome Proliferator-Activated Receptor Gamma Splicing Variants in the Buffalo Mammary Gland
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genomic Regions Associated with Resistance to Gastrointestinal Parasites in Australian Merino Sheep

by
Brenda Vera
1,
Elly A. Navajas
1,
Pablo Peraza
1,
Beatriz Carracelas
1,
Elize Van Lier
2,3 and
Gabriel Ciappesoni
1,*
1
Sistema Ganadero Extensivo, INIA Las Brujas, Canelones 90200, Uruguay
2
Departamento de Producción Animal y Pasturas, Facultad de Agronomía, Universidad de la República, Avda. Garzón 780, Montevideo 12900, Uruguay
3
Estación Experimental Facultad de Agronomía Salto, Salto 50000, Uruguay
*
Author to whom correspondence should be addressed.
Genes 2024, 15(7), 846; https://doi.org/10.3390/genes15070846
Submission received: 24 May 2024 / Revised: 22 June 2024 / Accepted: 24 June 2024 / Published: 27 June 2024
(This article belongs to the Special Issue Advances in Cattle, Sheep, and Goats Molecular Genetics and Breeding)

Abstract

:
The objective of this study was to identify genomic regions and genes associated with resistance to gastrointestinal nematodes in Australian Merino sheep in Uruguay, using the single-step GWAS methodology (ssGWAS), which is based on genomic estimated breeding values (GEBVs) obtained from a combination of pedigree, genomic, and phenotypic data. This methodology converts GEBVs into SNP effects. The analysis included 26,638 animals with fecal egg count (FEC) records obtained in two independent parasitic cycles (FEC1 and FEC2) and 1700 50K SNP genotypes. The comparison of genomic regions was based on genetic variances (gVar(%)) explained by non-overlap** regions of 20 SNPs. For FEC1 and FEC2, 18 and 22 genomic windows exceeded the significance threshold (gVar(%) ≥ 0.22%), respectively. The genomic regions with strong associations with FEC1 were located on chromosomes OAR 2, 6, 11, 21, and 25, and for FEC2 on OAR 5, 6, and 11. The proportion of genetic variance attributed to the top windows was 0.83% and 1.9% for FEC1 and FEC2, respectively. The 33 candidate genes shared between the two traits were subjected to enrichment analysis, revealing a marked enrichment in biological processes related to immune system functions. These results contribute to the understanding of the genetics underlying gastrointestinal parasite resistance and its implications for other productive and welfare traits in animal breeding programs.

Graphical Abstract

1. Introduction

The Australian Merino breed is a wool breed traditionally bred in Uruguay in extensive production systems, mainly located in the basalt region. Currently, this breed represents about 40% of the country’s sheep stock and is the main breed in fine and superfine wool production. One of the main problems affecting sheep production is infection by gastrointestinal parasites (GIPs) [1]. The decrease in production is a consequence of growth retardation, decreased weight gain, fleece weight and wool quality, and increased mortality [2,3], with Haemonchus contortus and Trichostrongylus colubriformis being the most prevalent parasites in Uruguay [4].
Given the problem of parasitism and the reported anthelmintic resistance [5,6,7], one alternative control method involves the breeding of animals genetically resistant to GIPs. The selection criterion used to assess resistance is the parasite fecal egg count (FEC), which is a moderately heritable trait [8] included in the National Genetic Evaluations (NGEs) in Uruguay. Currently, the NGE protocol includes the recording of two counts: the initial FEC (FEC1) shortly after weaning (7–9 months of age) and the subsequent FEC (FEC2) at 10–14 months of age [3].
Genome-wide association studies (GWASs) have been implemented in sheep populations to identify genes potentially associated with GIP resistance [9,10,11,12,13]. Unlike other methodologies that are limited to using genomic information exclusively from phenotyped animals, ssGWAS integrates genomic, pedigree, and phenotypic data from both genotyped and non-genotyped animals [14], thus considering population structure [15]. The ssGWAS procedure combines traditional pedigree relationships with those derived from genetic markers, and by the conversion of GEBVs to marker effects and weights [14].
The objectives of our study were as follows: (1) to estimate variance and heritability components for FEC1 and FEC2 traits in the Australian Merino sheep population in Uruguay; (2) to identify genomic regions and candidate genes associated with each trait by ssGWAS; and (3) to explore genes associated with FEC1 and FEC2, providing insight into the biological mechanisms underlying resistance to GIPs in sheep.

2. Materials and Methods

2.1. Natural Parasite Challenge

FEC determination was conducted according to the current protocol used for recording phenotypes in the NGEs [3]. In short, between 8 and 14 days after deworming at weaning, fecal samples were collected randomly from 15 to 20 lambs to assure the efficacy of the anthelmintic and that the animals started the evaluation period with 0 FECs. Every 14 to 21 days, 15 to 20 random lambs of each contemporary group (year, farm, sex) were sampled for FEC to evaluate the progress of the GIN infestation. When the mean of the FEC reached 500, and samples with 0 FECs represented less than 20% of the samples, all individuals were sampled for FEC1. After FEC1, lambs were again dewormed, and the process started over until all individuals were sampled for FEC2. All groups were managed under the same protocol, with lambs receiving oral drenching using proven anthelmintics such as Startec® (Zoetis, Auckland, New Zealand) and TritomNF® (Cibeles, Canelones, Uruguay).
The animals were probably exposed to multiple GIN species on pasture, as was observed in other studies [16]. The specific quantity of parasites ingested by each lamb could not be determined, but fecal sampling occurred during the warmer season (summer/fall) when environmental conditions were favorable for H. contortus and during the winter for Trichostrongylus sp. (Figure 1). In 2020, an analysis combining coproculture and (q)PCR was performed for the identification of GIN in grazing animals in Uruguay, and the results revealed a higher prevalence of H. contortus and Trichostrongylus sp. [17]. Previous studies report a prevalence of 43% of H. contortus and 38% of Trichostrongylus sp. [4]. Therefore, fecal samples were collected from two independent natural parasitic cycles separated by an anthelmintic treatment (Figure 1). Deworming of lambs at weaning and after FEC1 is essential to establish a baseline for comparison among individuals within a contemporary group.

2.2. Phenotypic Data

Samples were collected post-weaning; the mean age at recording (days) for FEC1 and FEC2 was 273 (±69) and 341 (±62), respectively. The mean time between each anthelmintic dose and sampling (days) for FEC1 (145 ± 64) and FEC2 (85 ± 40) exceeds the persistence period of anthelmintics in animals (≈15 days). A fecal sample was obtained from each individual directly from the rectum, and the modified McMaster technique with a sensitivity of 100 eggs per gram of feces was used to estimate the FEC [18]. Counts were transformed to natural logarithm, as described by Ciappesoni et al. [8], due to their non-normal distribution (LogFEC = Loge (FEC + 100)). In this study, we refer to logFEC1 and logFEC2 as FEC1 and FEC2, respectively.
A total of 26,638 animals born between 2001 and 2020 and belonging to 13 farms had FEC1 records. Among these, 18,971 animals also had FEC2 records (Table 1).

2.3. Genomic and Pedigree Data

Genomic DNA was extracted from blood samples, as described by Carracelas et al. [10]. A total of 1702 individuals were genotyped with the GeneSeek® Genomic Profiler™ (GGP, 43,705 SNPs) BeadChip (GeneSeek, Lincoln, NE, USA). Genomic data quality control was performed using preGSf90 (Aguilar et al., 2014). SNPs with a call rate below 90%, with minor allele frequency (MAF) less than 1%, monomorphic SNPs, and animals with call rate less than 90% were removed. Finally, 38,268 SNPs for 1697 sheep were used in the analysis.
The pedigree file was corrected using SeekParentF90 [19], which detects incompatibilities based on Mendelian conflict counts, as described in Wiggans et al. [20].

2.4. Statistical Analysis

Genetic parameters were estimated using methods based on pedigree relationships (BLUP) and pedigree–genomic models (ssGBLUP) [21]. A univariate model was conducted to estimate the variance components and heritabilities for FEC1 and FEC2 employing the AIREMLF90 software from the BLUPF90 family of programs [19]. Additionally, genetic correlations between FEC1 and FEC2 were estimated.
The following univariate model was used:
y = X b + Z u + e
where y is the vector of phenotypes for FEC1 or FEC2; X and Z are incidence matrices for fixed and random effects, respectively; b is the vector of fixed effects, including 494 contemporary groups (year of birth, sex, flock), dam age (three levels: 2, 3, and >4 years), type of birth (two levels: single or multiple), and lamb age at FEC1 or FEC2 recording as a covariate; u is the vector of random additive genetic effects; and e is the vector of residual effects.
In BLUP estimates, the random effects were modeled as u~N(0, A σ a 2 ) and e~N(0, I σ e 2 ), where A represents the numerator relationship matrix, I is the identity matrix, σ a 2 stands for the additive genetic variance, and σ e 2 is the residual variance. In the ssGBLUP model, the numerator relationship matrix ( A 1 ) utilized in BLUP is substituted with the H 1 matrix.
H 1 = A 1 + 0 0 0 G 1 A 22 1
where A 1 and A 22 1 are the inverse of the pedigree relationship matrix for all animals and for genotyped animals, respectively, and G 1 is the inverse of the genomic relationship matrix. The matrix G was constructed as described by Van Raden [22]:
G = Z D Z q
where Z is the SNP incidence matrix adjusted for allele frequencies, D is a weight matrix for SNP (initially D = I), and q is a weighting factor. The SNP effects and weighting factor were derived using an iterative process described by Wang et al. [14]. In this study, a single iteration was used as there was no significant change in SNP effects with additional iterations.
The percentage of genetic variance explained by region was calculated as follows:
V a r ( a i ) σ a 2 × 100 % = V a r Σ j = 1 20 Z j u ^ j σ a 2 × 100 %
where a i is the genetic value of the i-th region (20 contiguous SNPs), σ a 2 is the total genetic variance, Z j is the gene content vector of the j-th SNP for all individuals, and u ^ j is the effect of the j-th SNP marker within the i-th region [23].
Variance components and heritabilities were estimated by AIREMLF90. Heritability ( h 2 ) was calculated as h 2 = σ a 2 σ p 2 , and the total phenotypic variance ( σ p 2 ) was calculated as the sum of the additive genetic variance ( σ a 2 ) and the residual variance ( σ e 2 ).

2.5. Single-Step GWASs Analysis

ssGWAS is a two-step iterative procedure: (1) prediction of GEBVs using ssGBLUP, and (2) prediction of SNP effects based on GEBV. The detailed algorithm was described by Wang et al. [14].
The ssGWAS analysis was conducted independently for each trait and performed sequentially using RENUMF90 for general dataset preparation; PREGSF90 was used for quality control and generation of clean genotypes, and BLUPF90 and POSTGSF90 were used for the prediction of breeding values and SNP effects, respectively. These programs are part of the BLUPF90 software family, and for this study, the step-by-step tutorial reported by Masuda [24] was followed.

2.6. Identification of Candidate Genes and Functional Enrichment Analysis

Assuming that all windows explain the same proportion of genetic variance, the proportion of genetic variance explained by each of the 2519 1 Mb windows, including the 38,268 SNPs in the sheep genome, was 0.04%. Therefore, windows that explained at least 0.22% of the genetic variance, which is 5 times higher than expected (0.045 × 5 = 0.22%), were considered to contain putative QTL [25,26]. Regions representing 0.22% or more of the genetic variance σ a 2 were defined as significant regions. SNPs within these regions were identified and mapped onto the Oar_v3.1 sheep genome assembly using the Ensembl database [27]. A range of 5 kb upstream and downstream of the variant position was considered to identify candidate genes.
Functional enrichment analysis was conducted on the set of common candidate genes for both traits using DAVID (https://david.ncifcrf.gov/tools.jsp, accessed on 10 February 2024) [28]. Gene ontology (GO) terms with a p-value ≤ 0.05 were reported as significant terms.

3. Results

3.1. Variance Components and Heritabilities

The variance components and heritabilities for GIP resistance in Australian Merino sheep are presented in Table 2. The BLUP and ssGBLUP estimates of the heritabilities for FEC1 and FEC2 were close to 0.19. On the other hand, the genetic correlation between both traits was 0.88 (±0.03).

3.2. Genome-Wide Association Analysis

Figure 2 shows the Manhattan plots for GIP resistance traits. The genomic regions that explained the largest genetic variance for FEC1 were located on chromosomes 2, 6, 11, 21, and 25, and for FEC2, they were on chromosomes 5, 6, and 11. The top windows (most significant) explained 0.9% and 2% of the genetic variance for FEC1 and FEC2, respectively (Table 3).
A total of 18 and 22 windows with genetic variance greater than the significance threshold of 0.22 gVar (%) were identified, which included 316 and 376 SNPs related to FEC1 and FEC2 traits, respectively. Positional candidate genes close to SNPs (≤0.5 Mbps) were identified using the Ovis aries 3.1. reference genome map, and a total of 67 and 63 genes were mapped for FEC1 and FEC2, respectively. In total, 33 genes were shared between both traits (Figure 3). Details of the common positional candidate genes for the FEC1 and FEC2 traits are summarized in Table 4.

3.3. Enrichment Analysis

The ssGWAS results were complemented with a gene ontology (GO) enrichment analysis, which revealed 17 significantly enriched GO terms (p ≤ 0.05). Among these, ten were related to biological processes, two to molecular functions, and five to cellular components. The KEGG pathway enrichment analysis revealed three enriched metabolic pathways for the set of analyzed genes (Figure 4). Details of the enriched GO categories and the metabolic pathways involved are shown in Table 5.

4. Discussion

Variance component estimates obtained using the traditional pedigree-based approach (BLUP) were like those obtained using the ssGBLUP procedure, as well as the estimated heritability values (0.19 vs. 0.20). Medium-to-low heritabilities for resistance to GIPs in sheep [29,30] and for Australian Merino [8] have already been reported. Genetic parameter estimates obtained using ssGBLUP are known to be less biased and more accurate [31,32,33] since relationships between animals are better estimated [34]. In our case, the results were similar, and because there were no changes in the genetic base, the same additive variance is expected when including the genomic coefficients, as reported by Forni et al. [31].
The strong genetic correlation between FEC1 and FEC2 (0.88) suggests that they can be considered to be the same trait genetically, even when these traits were measured at different ages and correspond to two different parasitic cycles and seasons of the year (Figure 1), in which animals could have been exposed to different parasites. This high estimate is in agreement with other studies that explored the genetic association between FEC recorded at different ages in different breeds (i.e., 0.85 Romney; 0.82 Katahdin) [35,36]. A high genetic correlation of 0.74 between FEC by Strongyles sp. and Nematodirus sp. was also reported by Pacheco et al. [37].
On the other hand, ssGWAS revealed that the genomic regions reported as significant explain only 7 and 6% of the genetic variance for FEC1 and FEC2, respectively. These small variances suggest that resistance to GIPs is a polygenic trait with a large number of variants involved in the resistance mechanism [10,38,39]. Significant regions associated with GIP resistance have been previously reported on chromosomes 2, 6, 18, and 24 in several Australian sheep populations, including the Merino breed [40], and GIP-related QTL regions are also known [41].
In this study, seventeen positional candidate genes were identified on the OAR 2 for the FEC1 trait: AMER3, BMP1, CCAR2, CYFIP1, FAM160B2, HERC2, HR, NIPA1, NIPA2, PDLIM2, PIWIL2, POLR3D, PPP3CC, PTPN18, SORBS3, TUBGCP5, and XPO7. Among these, several are involved in the mechanisms of innate and adaptive immunity in mammals, such as HERC2 and CYFIP1, that are also involved in cytokine signaling [40,42]. The PDLIM2 gene has been associated in transcriptomic studies with the immune system and reproduction in sheep [43]. PTPN18 is involved in the B-cell receptor signaling pathway, being involved in differentiation, proliferation, and immunoglobulin (Ig) production [40], while other reports relate it to pigmentation in Merino sheep [44]. The SORBS3, PPP3CC, and PIWIL2 genes have been linked to growth and wool quality traits [45]. PPP3CC has been associated with heat tolerance in cattle [46], while PIWIL2 has been linked to reproductive traits in pigs [47]. In addition, these genes have been reported in selection signature studies, revealing their involvement in the parasite resistance of Slovakian sheep populations [48]. Al Kalaldeh et al. [40] also reported the association of the HERC2, NIPA1, NIPA2, CYFIP1, TUBGCP5, PTPN18, and AMER3 genes with GIP resistance in sheep and their involvement in immune system mechanisms. In addition, the CAST gene (OAR 5) was identified in the ssGWAS for FEC2 and it is known to have relevance to traits such as muscle production, carcass, and meat quality in sheep [49,50].
On the other hand, 12 candidate genes significantly associated with the FEC1 or FEC2 traits were identified on OAR 6: DCAF16, FAM184B, GPRIN3, HERC3, HERC5, IBSP, LAP3, LCORL, MED28, NCAPG, PIGY, and PYURF; all of them except LAP3 were detected by Al Kalaldeh et al. [40] in significant genomic regions for GIP resistance in sheep. Particularly, the FAM184B gene has also been reported to be strongly associated with body size and weight in livestock [51,52,53], milk production [54], and reproductive traits in sheep [55].
The MED28 gene is involved in milk production in ewes [45,54] and is also related to liveweight [56]. In addition, this gene has been linked to both pre- and postnatal body weights in ewes [51,56]. On the other hand, LAP3 also contributes to growth, milk production, and feed efficiency traits in sheep [57]. The GPRIN3 gene has been linked to prolificacy, litter size [58], and temperament in sheep [59].
Five candidate genes were identified on OAR 8: FAM120B, SMOC2, TBP, THBS2, WDR27, and WDR27. The THBS2 gene is related to growth and development traits [60] and high prolificacy in sheep [61], while the WDR27 gene has been associated with the cow milk fat trait [62].
On chromosome 11, a total of 43 candidate genes for FEC1 and FEC2 were identified, including ACADVL, ACAP1, ALOX12, ALOX15, ARRB2, ASGR1, BCL6B, C11H17orf49, C1QBP, AMTA2, CCDC42, CD68, CXCL16, DLG4, DNAH2, DVL2, INCA1, KIF1C, LOC101115381, LOC101121185, MED11, MINK1, MPDU1, MYH10, NDEL1, NUP88, ODF4, PELP1, PIK3R5, PLD2, RABEP1, RPAIN, SENP3, SLC16A13, SLC25A35, SPAG7, STX8, TNFSF12, USP43, WRAP53, WSCD1, ZFP3, and ZMYND15. The ALOX15 gene is known to play a role in H. contortus infection in sheep [63]. Other studies have reported QTLs for T. colubriformis egg count in the Merino breed in this genomic region [64], as well as QTLs for height, weight, and parasite resistance in the region between 32.13 and 32.19 Mbp [65]. In addition, selection signatures for resistance to H. contortus in sheep and goats have been reported by Estrada-Reyes et al. [66], while the CTNNA gene, located on chromosome 25, has been significantly associated with FEC1 and FEC2 traits and is related to growth traits [67] and brucellosis resistance in sheep [68].
The GO analysis (Figure 4) revealed multiple GO terms related to host defense mechanisms against pathogens, including the lipoxin A4 biosynthetic process (GO:2001303). Lipoxins have been reported to be endogenous anti-inflammatory molecules involved in reducing excessive tissue damage and chronic inflammation [69]. These lipoxins can be synthesized from arachidonic acid, which justifies the over-representation of biological processes such as the lipoxygenase pathway (GO:0019372) and arachidonic acid metabolic process (GO:0019369), as well as arachidonate 15-lipoxygenase activity and linoleate 13S-lipoxygenase activity molecular functions (GO:0050473, GO:001616).
Cytokines, such as IL-12, are critical for host resistance to many pathogens but can also be detrimental when expressed in an uncontrolled manner [70,71], so it makes biological sense that the GO term the negative regulation of interleukin-12 production (GO:0032695) is over-represented. The role of lipoxins in mediating the immune response has been studied against parasitic pathogens [72] and in other diseases [69]. Another GO term related to host defense against pathogens is the GO term the linoleic acid metabolic process (GO:0043651). The activation of linoleic acid metabolism in macrophages in bacteria promotes pathogen killing [73]. In that sense, one of the most over-represented genes is the ALOX15 gene, which is related to the oxidation of arachidonic acid and the production of anti-inflammatory lipoxins [74], and it was identified as a differentially expressed gene associated with sheep resistance to the nematode Teladorsagia circumcincta [75].
As with all infections, parasites cause inflammation, involving the mobilization, proliferation, and recruitment of leukocytes to the affected area. This trafficking of immune cells and the effector functions of these cells can serve to control the pathogen or exacerbate the pathology [76]. In this regard, one of the most over-represented metabolic pathways is the oas04062/chemokine signaling pathway. Two of the functions of chemokines are to attract immune cells to sites of inflammation [77] and to guide the migration of neurons and other migratory cells [78].
In summary, the enrichment analysis shows that the candidate genes enriched biological processes and molecular functions related to the metabolism of linoleic acid, which is a metabolic precursor of arachidonic acid. Free arachidonic acid and its metabolites promote and modulate the type 2 immune response, playing a crucial role in GIP resistance through the action of eosinophils, basophils, and mast cells [79].

5. Conclusions

In this study, genetic parameters were estimated to evaluate resistance to GIPs in two independent parasite cycles, each separated by anthelmintic treatment. A total of 18 and 22 genomic regions were identified that showed a significant association with FEC1 and FEC2, respectively. We report positional candidate genes for both cycles using ssGWAS in Australian Merino sheep, some of which are novel for these traits. Our study reveals a set of candidate genes that share mechanisms related to immune response, body size, and weight, as well as genes associated with reproductive traits. In summary, our findings provide a basis for future genomic research and could contribute significantly to breeding programs.

Author Contributions

B.V.: investigation, data curation, formal analysis, writing—original draft, writing—review and editing; B.C.: resources, methodology, writing—review and editing; P.P.: resources, methodology, writing—review and editing; E.A.N.: conceptualization, methodology, investigation, writing—review and editing, project administration, funding acquisition; E.V.L.: writing—review and editing, funding acquisition; G.C.: conceptualization, writing—review and editing, project administration, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Instituto Nacional de Investigación Agropecuaria (INIA_CL_40 and INIA_CL_38), CSIC_I+D_2018_287 (Comisión Sectorial de Investigación Científica, CSIC, Universidad de la República), and the European Union’s Horizon 2020 research and innovation program (SMARTER, agreement no. 772787).

Institutional Review Board Statement

The animal study protocol was approved by the Animal Ethics Committee of INIA (approval number INIA_2018.2, July 2018) and approved by the Honorary Commission for Animal Experimentation (CHEA, protocols N° 021130-006307-11 and 020300-000653-18-ID, CHEA 701, Universidad de la República).

Informed Consent Statement

Not applicable.

Data Availability Statement

Restrictions apply to the availability of these data. Data were obtained by INIA and are available from the authors with INIA’s permission.

Acknowledgments

We thank Ignacio Aguilar for his help using BLUPF90 programs, the Australian Merino Breeders Association of Uruguay and Estación Experimental Facultad de Agronomía Salto, and the INIA staff who collaborated in collecting data for this research, especially Carlos Monzalvo for managing the phenotypic database.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zajac, A.M. Gastrointestinal Nematodes of Small Ruminants: Life Cycle, Anthelmintics, and Diagnosis. Vet. Clin. N. Am. Food Anim. Pract. 2006, 22, 529–541. [Google Scholar] [CrossRef] [PubMed]
  2. Zajac, A.M.; Garza, J. Biology, Epidemiology, and Control of Gastrointestinal Nematodes of Small Ruminants. Vet. Clin. N. Am. Food Anim. Pract. 2020, 36, 73–87. [Google Scholar] [CrossRef] [PubMed]
  3. Castells, D. Evaluación de Resistencia de Ovinos Corriedale a Los Nematodos En Uruguay: Heredabilidad y Correlaciones Genéticas El Recuento De Huevos De Nematodos Características Productivas. Master’s Thesis, UDELAR, Montevideo, Uruguay, 2008. [Google Scholar]
  4. Nari, A.; Cardozo, H.; Berdié, J.; Canábez, F.; Bawden, R. Dinámica de población para nematodes gastrointestinales de ovinos en Uruguay: Trabajo asistido por UNDP y FAO. Vet. Montev. 1977, 14, 11–23. [Google Scholar]
  5. Mederos, A.E.; Ramos, Z.; Banchero, G.E. First Report of Monepantel Haemonchus Contortus Resistance on Sheep Farms in Uruguay. Parasit. Vectors 2014, 7, 598. [Google Scholar] [CrossRef] [PubMed]
  6. Roeber, F.; Jex, A.R.; Gasser, R.B. Impact of Gastrointestinal Parasitic Nematodes of Sheep, and the Role of Advanced Molecular Tools for Exploring Epidemiology and Drug Resistance—An Australian Perspective. Parasit. Vectors 2013, 6, 153. [Google Scholar] [CrossRef] [PubMed]
  7. Waller, P.J.; Echevarria, F.; Eddi, C.; Maciel, S.; Nari, A.; Hansen, J.W. The Prevalence of Anthelmintic Resistance in Nematode Parasites of Sheep in Southern Latin America: General Overview. Vet. Parasitol. 1996, 62, 181–187. [Google Scholar] [CrossRef] [PubMed]
  8. Ciappesoni, G.; Goldberg, V.; Gimeno, D. Estimates of Genetic Parameters for Worm Resistance, Wool and Growth Traits in Merino Sheep of Uruguay. Livest. Sci. 2013, 157, 65–74. [Google Scholar] [CrossRef]
  9. Arzik, Y.; Kizilaslan, M.; White, S.N.; Piel, L.M.W.; Çınar, M.U. Genomic Analysis of Gastrointestinal Parasite Resistance in Akkaraman Sheep. Genes 2022, 13, 2177. [Google Scholar] [CrossRef] [PubMed]
  10. Carracelas, B.; Navajas, E.A.; Vera, B.; Ciappesoni, G. Genome-Wide Association Study of Parasite Resistance to Gastrointestinal Nematodes in Corriedale Sheep. Genes 2022, 13, 1548. [Google Scholar] [CrossRef]
  11. Estrada-Reyes, Z.M.; Ogunade, I.M.; Pech-Cervantes, A.A.; Terrill, T.H. Copy Number Variant-Based Genome Wide Association Study Reveals Immune-Related Genes Associated with Parasite Resistance in a Heritage Sheep Breed from the United States. Parasite Immunol. 2022, 44, e12943. [Google Scholar] [CrossRef]
  12. Pacheco, A.; Conington, J.; Corripio-Miyar, Y.; Frew, D.; Banos, G.; McNeilly, T.N. Genetic Profile of Adaptive Immune Traits and Relationships with Parasite Resistance and Productivity in Scottish Blackface Sheep. Animal 2024, 18, 101061. [Google Scholar] [CrossRef] [PubMed]
  13. Thorne, J.W.; Redden, R.; Bowdridge, S.A.; Becker, G.M.; Stegemiller, M.R.; Murdoch, B.M. Genome-Wide Analysis of Sheep Artificially or Naturally Infected with Gastrointestinal Nematodes. Genes 2023, 14, 1342. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, H.; Misztal, I.; Aguilar, I.; Legarra, A.; Muir, W.M. Genome-Wide Association Map** Including Phenotypes from Relatives without Genotypes. Genet. Res. 2012, 94, 73–83. [Google Scholar] [CrossRef] [PubMed]
  15. Mancin, E.; Lourenco, D.; Bermann, M.; Mantovani, R.; Misztal, I. Accounting for Population Structure and Phenotypes from Relatives in Association Map** for Farm Animals: A Simulation Study. Front. Genet. 2021, 12, 642065. [Google Scholar] [CrossRef] [PubMed]
  16. Pereira, D.; Castells, D.; Deschenaux, H. Infectividad del campo natural contaminado con huevos de Haemonchus contortus en cuatro estaciones del año. Prod. Ovina 2006, 18, 57–65. [Google Scholar]
  17. Pimentel Barreto, S. Puesta a Punto y Validación de Metodologías Basadas en ADN para Diagnóstico de Nemátodos Gastrointestinales de Ovinos en Uruguay. Master’s Thesis, Udelar. FV, Montevideo, Uruguay, 2020. [Google Scholar]
  18. Whitlock, H.V. Some Modifications of the McMaster Helminth Egg-Counting Technique and Apparatus. J. Sci. Ind. Res. 1948, 21, 177–180. [Google Scholar]
  19. Misztal, I.; Tsuruta, S.; Lourenco, D.; Masuda, Y.; Aguilar, I. BLUPF90 Family of Programs; University of Georgia: Athens, GA, USA. Available online: https://nce.ads.uga.edu/wiki/doku.php?id=start (accessed on 22 February 2023).
  20. Wiggans, G.R.; VanRaden, P.M.; Bacheller, L.R.; Tooker, M.E.; Hutchison, J.L.; Cooper, T.A.; Sonstegard, T.S. Selection and Management of DNA Markers for Use in Genomic Evaluation. J. Dairy Sci. 2010, 93, 2287–2292. [Google Scholar] [CrossRef] [PubMed]
  21. Aguilar, I.; Misztal, I.; Johnson, D.L.; Legarra, A.; Tsuruta, S.; Lawlor, T.J. Hot Topic: A Unified Approach to Utilize Phenotypic, Full Pedigree, and Genomic Information for Genetic Evaluation of Holstein Final Score1. J. Dairy Sci. 2010, 93, 743–752. [Google Scholar] [CrossRef]
  22. VanRaden, P.M. Efficient Methods to Compute Genomic Predictions. J. Dairy Sci. 2008, 91, 4414–4423. [Google Scholar] [CrossRef]
  23. Wang, H.; Misztal, I.; Aguilar, I.; Legarra, A.; Fernando, R.L.; Vitezica, Z.; Okimoto, R.; Wing, T.; Hawken, R.; Muir, W.M. Genome-Wide Association Map** Including Phenotypes from Relatives without Genotypes in a Single-Step (ssGWAS) for 6-Week Body Weight in Broiler Chickens. Front. Genet. 2014, 5, 87369. [Google Scholar] [CrossRef]
  24. Yutaka Masuda. Available online: https://masuday.github.io/ (accessed on 12 February 2024).
  25. Sollero, B.P.; Junqueira, V.S.; Gomes, C.C.G.; Caetano, A.R.; Cardoso, F.F. Tag SNP Selection for Prediction of Tick Resistance in Brazilian Braford and Hereford Cattle Breeds Using Bayesian Methods. Genet. Sel. Evol. 2017, 49, 49. [Google Scholar] [CrossRef] [PubMed]
  26. Onteru, S.K.; Gorbach, D.M.; Young, J.M.; Garrick, D.J.; Dekkers, J.C.M.; Rothschild, M.F. Whole Genome Association Studies of Residual Feed Intake and Related Traits in the Pig. PLoS ONE 2013, 8, e61756. [Google Scholar] [CrossRef] [PubMed]
  27. International Sheep Genomics Consortium; Archibald, A.L.; Cockett, N.E.; Dalrymple, B.P.; Faraut, T.; Kijas, J.W.; Maddox, J.F.; McEwan, J.C.; Hutton Oddy, V.; Raadsma, H.W.; et al. The Sheep Genome Reference Sequence: A Work in Progress. Anim. Genet. 2010, 41, 449–453. [Google Scholar] [CrossRef]
  28. Sherman, B.T.; Hao, M.; Qiu, J.; Jiao, X.; Baseler, M.W.; Lane, H.C.; Imamichi, T.; Chang, W. DAVID: A Web Server for Functional Enrichment Analysis and Functional Annotation of Gene Lists (2021 Update). Nucleic Acids Res. 2022, 50, W216–W221. [Google Scholar] [CrossRef] [PubMed]
  29. Gutiérrez-Gil, B.; Pérez, J.; de la Fuente, L.F.; Meana, A.; Martínez-Valladares, M.; San Primitivo, F.; Rojo-Vázquez, F.A.; Arranz, J.J. Genetic Parameters for Resistance to Trichostrongylid Infection in Dairy Sheep. Animal 2010, 4, 505–512. [Google Scholar] [CrossRef] [PubMed]
  30. Safari, E.; Fogarty, N.M.; Gilmour, A.R. A Review of Genetic Parameter Estimates for Wool, Growth, Meat and Reproduction Traits in Sheep. Livest. Prod. Sci. 2005, 92, 271–289. [Google Scholar] [CrossRef]
  31. Forni, S.; Aguilar, I.; Misztal, I. Different Genomic Relationship Matrices for Single-Step Analysis Using Phenotypic, Pedigree and Genomic Information. Genet. Sel. Evol. 2011, 43, 1. [Google Scholar] [CrossRef] [PubMed]
  32. Gordo, D.G.M.; Espigolan, R.; Tonussi, R.L.; Júnior, G.A.F.; Bresolin, T.; Magalhães, A.F.B.; Feitosa, F.L.; Baldi, F.; Carvalheiro, R.; Tonhati, H.; et al. Genetic Parameter Estimates for Carcass Traits and Visual Scores Including or Not Genomic Information1. J. Anim. Sci. 2016, 94, 1821–1826. [Google Scholar] [CrossRef] [PubMed]
  33. Tsuruta, S.; Misztal, I.; Lourenco, D.A.L.; Lawlor, T.J. Assigning Unknown Parent Groups to Reduce Bias in Genomic Evaluations of Final Score in US Holsteins. J. Dairy Sci. 2014, 97, 5814–5821. [Google Scholar] [CrossRef]
  34. Lourenco, D.A.L.; Tsuruta, S.; Fragomeni, B.O.; Masuda, Y.; Aguilar, I.; Legarra, A.; Bertrand, J.K.; Amen, T.S.; Wang, L.; Moser, D.W.; et al. Genetic Evaluation Using Single-Step Genomic Best Linear Unbiased Predictor in American Angus1. J. Anim. Sci. 2015, 93, 2653–2662. [Google Scholar] [CrossRef]
  35. Morris, C.; Wheeler, M. Correlated Responses Following Genetic Selection to Change Faecal Egg Count in Romney. Proc. N. Z. Soc. Anim. Prod. 2010, 70, 229–234. [Google Scholar]
  36. Ngere, L.; Burke, J.M.; Morgan, J.L.M.; Miller, J.E.; Notter, D.R. Genetic Parameters for Fecal Egg Counts and Their Relationship with Body Weights in Katahdin Lambs. J. Anim. Sci. 2018, 96, 1590–1599. [Google Scholar] [CrossRef] [PubMed]
  37. Pacheco, A.; McNeilly, T.N.; Banos, G.; Conington, J. Genetic Parameters of Animal Traits Associated with Coccidian and Nematode Parasite Load and Growth in Scottish Blackface Sheep. Animal 2021, 15, 100185. [Google Scholar] [CrossRef] [PubMed]
  38. Kemper, K.E.; Emery, D.L.; Bishop, S.C.; Oddy, H.; Hayes, B.J.; Dominik, S.; Henshall, J.M.; Goddard, M.E. The Distribution of SNP Marker Effects for Faecal Worm Egg Count in Sheep, and the Feasibility of Using These Markers to Predict Genetic Merit for Resistance to Worm Infections. Genet. Res. 2011, 93, 203–219. [Google Scholar] [CrossRef] [PubMed]
  39. Riggio, V.; Matika, O.; Pong-Wong, R.; Stear, M.J.; Bishop, S.C. Genome-Wide Association and Regional Heritability Map** to Identify Loci Underlying Variation in Nematode Resistance and Body Weight in Scottish Blackface Lambs. Heredity 2013, 110, 420–429. [Google Scholar] [CrossRef] [PubMed]
  40. Al Kalaldeh, M.; Gibson, J.; Lee, S.H.; Gondro, C.; van der Werf, J.H.J. Detection of Genomic Regions Underlying Resistance to Gastrointestinal Parasites in Australian Sheep. Genet. Sel. Evol. GSE 2019, 51, 37. [Google Scholar] [CrossRef] [PubMed]
  41. Cunha, S.M.F.; Lam, S.; Mallard, B.; Karrow, N.A.; Cánovas, Á. Genomic Regions Associated with Resistance to Gastrointestinal Nematode Parasites in Sheep—A Review. Genes 2024, 15, 187. [Google Scholar] [CrossRef] [PubMed]
  42. Abied, A.; Bagadi, A.; Bordbar, F.; Pu, Y.; Augustino, S.M.A.; Xue, X.; ** Identifies Genes with Pleiotropic Effects on Body Composition. BMC Genom. 2016, 17, 224. [Google Scholar] [CrossRef]
  43. Yuan, Z.; Sunduimijid, B.; ** Reveals Signatures of Selection Related to Acclimation and Economically Important Traits in 15 Local Sheep Breeds from Russia. BMC Genom. 2019, 20, 294. [Google Scholar] [CrossRef] [PubMed]
  44. Tao, L.; He, X.; Jiang, Y.; Liu, Y.; Ouyang, Y.; Shen, Y.; Hong, Q.; Chu, M. Genome-Wide Analyses Reveal Genetic Convergence of Prolificacy between Goats and Sheep. Genes 2021, 12, 480. [Google Scholar] [CrossRef] [PubMed]
  45. Romaniuk, E.; Vera, B.; Peraza, P.; Ciappesoni, G.; Damián, J.P.; Van Lier, E. Identification of Candidate Genes and Pathways Linked to the Temperament Trait in Sheep. Genes 2024, 15, 229. [Google Scholar] [CrossRef] [PubMed]
  46. Zhu, M.; Yang, Y.; Yang, H.; Zhao, Z.; Zhang, H.; Blair, H.T.; Zheng, W.; Wang, M.; Fang, C.; Yu, Q.; et al. Whole-Genome Resequencing of the Native Sheep Provides Insights into the Microevolution and Identifies Genes Associated with Reproduction Traits. BMC Genom. 2023, 24, 392. [Google Scholar] [CrossRef] [PubMed]
  47. Chen, H.Y.; Shen, H.; Jia, B.; Zhang, Y.S.; Wang, X.H.; Zeng, X.C. Differential Gene Expression in Ovaries of Qira Black Sheep and Hetian Sheep Using RNA-Seq Technique. PLoS ONE 2015, 10, e0120170. [Google Scholar] [CrossRef]
  48. Peters, S.O.; Kızılkaya, K.; Ibeagha-Awemu, E.M.; Sinecen, M.; Zhao, X. Comparative Accuracies of Genetic Values Predicted for Economically Important Milk Traits, Genome-Wide Association, and Linkage Disequilibrium Patterns of Canadian Holstein Cows. J. Dairy Sci. 2021, 104, 1900–1916. [Google Scholar] [CrossRef]
  49. Guo, Z.; González, J.F.; Hernandez, J.N.; McNeilly, T.N.; Corripio-Miyar, Y.; Frew, D.; Morrison, T.; Yu, P.; Li, R.W. Possible Mechanisms of Host Resistance to Haemonchus Contortus Infection in Sheep Breeds Native to the Canary Islands. Sci. Rep. 2016, 6, 26200. [Google Scholar] [CrossRef]
  50. Beh, K.J.; Hulme, D.J.; Callaghan, M.J.; Leish, Z.; Lenane, I.; Windon, R.G.; Maddox, J.F. A Genome Scan for Quantitative Trait Loci Affecting Resistance to Trichostrongylus Colubriformis in Sheep. Anim. Genet. 2002, 33, 97–106. [Google Scholar] [CrossRef]
  51. Niciura, S.C.M.; Benavides, M.V.; Okino, C.H.; Ibelli, A.M.G.; Minho, A.P.; Esteves, S.N.; de Souza Chagas, A.C. Genome-Wide Association Study for Haemonchus Contortus Resistance in Morada Nova Sheep. Pathogens 2022, 11, 939. [Google Scholar] [CrossRef] [PubMed]
  52. Estrada-Reyes, Z.M.; Tsukahara, Y.; Amadeu, R.R.; Goetsch, A.L.; Gipson, T.A.; Sahlu, T.; Puchala, R.; Wang, Z.; Hart, S.P.; Mateescu, R.G. Signatures of Selection for Resistance to Haemonchus Contortus in Sheep and Goats. BMC Genom. 2019, 20, 735. [Google Scholar] [CrossRef] [PubMed]
  53. Zhao, L.; Li, F.; Yuan, L.; Zhang, X.; Zhang, D.; Li, X.; Zhang, Y.; Zhao, Y.; Song, Q.; Wang, J.; et al. Expression of Ovine CTNNA3 and CAP2 Genes and Their Association with Growth Traits. Gene 2022, 807, 145949. [Google Scholar] [CrossRef] [PubMed]
  54. Li, X.; Wu, Q.; Zhang, X.; Li, C.; Zhang, D.; Li, G.; Zhang, Y.; Zhao, Y.; Shi, Z.; Wang, W.; et al. Whole-Genome Resequencing to Study Brucellosis Susceptibility in Sheep. Front. Genet. 2021, 12, 653927. [Google Scholar] [CrossRef] [PubMed]
  55. Chandrasekharan, J.A.; Sharma-Walia, N. Lipoxins: Nature’s Way to Resolve Inflammation. J. Inflamm. Res. 2015, 8, 181–192. [Google Scholar] [CrossRef] [PubMed]
  56. Aliberti, J.; Serhan, C.; Sher, A. Parasite-Induced Lipoxin A4 Is an Endogenous Regulator of IL-12 Production and Immunopathology in Toxoplasma Gondii Infection. J. Exp. Med. 2002, 196, 1253–1262. [Google Scholar] [CrossRef] [PubMed]
  57. Eyles, J.L.; Metcalf, D.; Grusby, M.J.; Hilton, D.J.; Starr, R. Regulación Negativa de La Señalización de Interleucina-12 Por El Supresor de La Señalización de Citocina-1 *. J. Biol. Chem. 2002, 277, 43735–43740. [Google Scholar] [CrossRef]
  58. Wei, F.; Gong, W.; Wang, J.; Yang, Y.; Liu, J.; Wang, Y.; Cao, J. Role of the Lipoxin A4 Receptor in the Development of Neutrophil Extracellular Traps in Leishmania Infantum Infection. Parasit. Vectors 2019, 12, 275. [Google Scholar] [CrossRef]
  59. Yan, B.; Fung, K.; Ye, S.; Lai, P.-M.; **n Wei, Y.; Sze, K.-H.; Yang, D.; Gao, P.; Yi-Tsun Kao, R. Linoleic Acid Metabolism Activation in Macrophages Promotes the Clearing of Intracellular Staphylococcus Aureus. Chem. Sci. 2022, 13, 12445–12460. [Google Scholar] [CrossRef]
  60. Samuelsson, B.; Dahlén, S.E.; Lindgren, J.A.; Rouzer, C.A.; Serhan, C.N. Leukotrienes and Lipoxins: Structures, Biosynthesis, and Biological Effects. Science 1987, 237, 1171–1176. [Google Scholar] [CrossRef]
  61. Wilkie, H.; Xu, S.; Gossner, A.; Hopkins, J. Variable Exon Usage of Differentially-Expressed Genes Associated with Resistance of Sheep to Teladorsagia Circumcincta. Vet. Parasitol. 2015, 212, 206–213. [Google Scholar] [CrossRef] [PubMed]
  62. McGovern, K.E.; Wilson, E.H. Role of Chemokines and Trafficking of Immune Cells in Parasitic Infections. Curr. Immunol. Rev. 2013, 9, 157–168. [Google Scholar] [CrossRef] [PubMed]
  63. Oppenheim, J.J.; Zachariae, C.O.C.; Mukaida, N.; Matsushima, K. Properties of the Novel Proinflammatory Supergene “Intercrine” Cytokine Family. Annu. Rev. Immunol. 1991, 9, 617–648. [Google Scholar] [CrossRef] [PubMed]
  64. Wang, J.; Knaut, H. Chemokine Signaling in Development and Disease. Development 2014, 141, 4199–4205. [Google Scholar] [CrossRef]
  65. Tallima, H.; El Ridi, R. Arachidonic Acid: Physiological Roles and Potential Health Benefits—A Review. J. Adv. Res. 2017, 11, 33–41. [Google Scholar] [CrossRef]
Figure 1. Sampling scheme for fecal egg counts (FECs) in two independent parasitic cycles and in different seasons.
Figure 1. Sampling scheme for fecal egg counts (FECs) in two independent parasitic cycles and in different seasons.
Genes 15 00846 g001
Figure 2. Manhattan plots depict the genetic variance explained (%) by 20 adjacent SNP windows for FEC1 (a) and FEC2 (b) in Australian Merino sheep. Each dot represents a window, with the percentage of additive genetic variance explained by each window. The horizontal line indicates the suggestive threshold of 0.22 of the gVar (%).
Figure 2. Manhattan plots depict the genetic variance explained (%) by 20 adjacent SNP windows for FEC1 (a) and FEC2 (b) in Australian Merino sheep. Each dot represents a window, with the percentage of additive genetic variance explained by each window. The horizontal line indicates the suggestive threshold of 0.22 of the gVar (%).
Genes 15 00846 g002
Figure 3. List of positional candidate genes identified in genomic regions that explain > 0.22 gVar(%) for FEC1 and FEC2 (a). The black dot represents that this gene was identified in the analysis for each trait. The bar plot shows the number of candidate genes in common for both traits at the intersection (b).
Figure 3. List of positional candidate genes identified in genomic regions that explain > 0.22 gVar(%) for FEC1 and FEC2 (a). The black dot represents that this gene was identified in the analysis for each trait. The bar plot shows the number of candidate genes in common for both traits at the intersection (b).
Genes 15 00846 g003
Figure 4. Gene ontology (GO) term enrichment analysis of genes in common associated with FEC1 and FEC2 traits in Australian Merino sheep. Categories included biological process (BP), cellular component (CC), and molecular function (MF) on the x-axis the -log10 of the adjusted p-value (<0.05) and on the y-axis the GO term.
Figure 4. Gene ontology (GO) term enrichment analysis of genes in common associated with FEC1 and FEC2 traits in Australian Merino sheep. Categories included biological process (BP), cellular component (CC), and molecular function (MF) on the x-axis the -log10 of the adjusted p-value (<0.05) and on the y-axis the GO term.
Genes 15 00846 g004
Table 1. Descriptive statistics for fecal egg counts (LogFEC1 and LogFEC2).
Table 1. Descriptive statistics for fecal egg counts (LogFEC1 and LogFEC2).
TraitNMeanSD aMin bMax c
LogFEC126,6386.631.144.6010.51
LogFEC218,9716.641.134.6010.87
a SD: standard deviation; b Min: minimum; c Max: maximum.
Table 2. Additive genetic variances, residuals, and heritability for FEC1 and FEC2 in Australian Merino sheep.
Table 2. Additive genetic variances, residuals, and heritability for FEC1 and FEC2 in Australian Merino sheep.
TraitMethod σ a 2 σ e 2 h2
FEC1BLUP0.15 ± 0.010.66 ± 0.010.19 ± 0.01
SSGBLUP0.16 ± 0.010.66 ± 0.010.20 ± 0.01
FEC2BLUP0.15 ± 0.010.66 ± 0.010.19 ± 0.02
SSGBLUP0.16 ± 0.010.66 ± 0.010.19 ± 0.02
σ a 2 = additive genetic variance; σ e 2 = residual variance; h 2 = heritability.
Table 3. Chromosome, location, proportion of genetic variance, and candidate genes within the top 10 windows associated with the FEC1 and FEC2 traits in Australian Merino sheep.
Table 3. Chromosome, location, proportion of genetic variance, and candidate genes within the top 10 windows associated with the FEC1 and FEC2 traits in Australian Merino sheep.
TraitChr agVar(%) bWindows Bounds (pb)Candidate Genes
FEC160.8335,429,074–35,554,875GPRIN3
60.7337,072,532–37,134,706LAP3, MED28, FAM184B
110.6126,119,722–26,407,016MINK1, PLD2, ZMYND15, CXCL16, MED11, ARRB2, PELP1, ALOX15
110.5727,002,322–27,807,356DNAH2, LOC101121185, SLC25A35, ODF4, NDEL1, MYH10
250.4921,324,145–22,654,093CTNNA3
110.3625,298,833–25,633,178WSCD1
60.3036,066,911–36,197,551HERC3, PYURF, PIGY, HERC5
210.3042,654,067–42,734,621SYVN1, SPDYC, CAPN1
20.29112,528,931–113,355,547HERC2, NIPA1, NIPA2, CYFIP1, TUBGCP5, PTPN18, AMER3
110.2827,811,348–28,475,634CCDC42, PIK3R5, STX8, USP43
FEC2111.9426,119,722–26,407,016MINK1, PLD2, ZMYND15, CXCL16, MED11, ARRB2, PELP1, ALOX15
51.1393,463,383–93,482,724CAST
110.9027,002,322–27,807,356DNAH2, LOC101121185, SLC25A35, ODF4, NDEL1, MYH10
110.7727,811,348–28,475,634CCDC42, PIK3R5, STX8, USP43
110.7625,298,833–25,633,178WSCD1
50.6893,437,720–93,463,383CAST
60.6437,072,532–37,134,706LAP3, MED28, FAM184B
60.5335,429,074–35,554,875GPRIN3
110.5225,744,499–26,098,771C1QBP, RPAIN, NUP88, RABEP1, ZFP3, KIF1C, INCA1, CAMTA2, SPAG7
60.5137,483,582–37,665,116-
a Chr = chromosome number. b gVar(%) = proportion of genetic variance represented by each region comprising 20 SNPs.
Table 4. Positional candidate genes in common for FEC1 and FEC2.
Table 4. Positional candidate genes in common for FEC1 and FEC2.
ChrPos (bp)Candidate Genes
635,511,497–37,257,065FAM184B, MED28, LAP3, GPRIN3
1125,298,833–28,475,634USP43, STX8, PIK3R5, CCDC42, MYH10, NDEL1, ODF4, SLC25A35, LOC101121185, DNAH2, ALOX15, PELP1, ARRB2, CXCL16, MED11, ZMYND15, PLD2, MINK1, CAMTA2, SPAG7, INCA1, KIF1C, ZFP3, RABEP1, NUP88, C1QBP, RPAIN, WSCD1
2522,143,113–22,654,093CTNNA3
Table 5. Gene ontology (GO) terms such as biological processes, molecular functions, and KEGG pathways of candidate genes associated with FEC1 and FEC2 traits in Australian Merino sheep.
Table 5. Gene ontology (GO) terms such as biological processes, molecular functions, and KEGG pathways of candidate genes associated with FEC1 and FEC2 traits in Australian Merino sheep.
CategoryTermGenesp-Value−log10 (p-Value)
GOTERM_BPGO:2001303~lipoxin A4 biosynthetic processALOX15, LOC1011211850.0052.289
GOTERM_BPGO:0001764~neuron migrationNDEL1, ENSOARG00000002682, MYH100.0072.140
GOTERM_BPGO:0019372~lipoxygenase pathwayALOX15, LOC1011211850.0101.989
GOTERM_BPGO:0043651~linoleic acid metabolic processALOX15, LOC1011211850.0101.989
GOTERM_BPGO:0007097~nuclear migrationENSOARG00000002682, MYH100.0141.865
GOTERM_BPGO:0032695~negative regulation of interleukin-12 productionARRB2, C1QBP0.0201.690
GOTERM_BPGO:0019369~arachidonic acid metabolic processALOX15, LOC1011211850.0201.690
GOTERM_BPGO:0030334~regulation of cell migrationMINK1, ENSOARG000000026820.0751.128
GOTERM_BPGO:0007286~spermatid developmentCCDC42, ZMYND150.0821.084
GOTERM_BPGO:0007018~microtubule-based movementDNAH2, KIF1C0.0931.029
GOTERM_CCGO:0070847~core mediator complexMED11, MED280.0201.690
GOTERM_CCGO:0005737~cytoplasmNDEL1, LOC101121185, CTNNA3, ENSOARG00000002682, PELP1, INCA1, ZMYND15, ENSOARG000000070660.0371.427
GOTERM_CCGO:0030139~endocytic vesicleRABEP1, ARRB20.0491.313
GOTERM_CCGO:0005819~spindleNDEL1, MYH100.0811.092
GOTERM_CCGO:0005829~cytosolALOX15, LOC101121185, PIK3R5, STX8, LAP3, ARRB2, C1QBP, MYH100.0821.086
GOTERM_MFGO:0050473~arachidonate 15-lipoxygenase activityALOX15, LOC1011211850.0052.301
GOTERM_MFGO:0016165~linoleate 13S-lipoxygenase activityALOX15, LOC1011211850.0072.176
KEGG_PATHWAYoas04062/Chemokine signaling pathwayPIK3R5, ARRB2, CXCL160.0331.486
KEGG_PATHWAYoas04814/Motor proteinsDNAH2, MYH10, KIF1C0.0331.478
KEGG_PATHWAYoas04144/EndocytosisRABEP1, PLD2, ARRB20.0551.260
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vera, B.; Navajas, E.A.; Peraza, P.; Carracelas, B.; Van Lier, E.; Ciappesoni, G. Genomic Regions Associated with Resistance to Gastrointestinal Parasites in Australian Merino Sheep. Genes 2024, 15, 846. https://doi.org/10.3390/genes15070846

AMA Style

Vera B, Navajas EA, Peraza P, Carracelas B, Van Lier E, Ciappesoni G. Genomic Regions Associated with Resistance to Gastrointestinal Parasites in Australian Merino Sheep. Genes. 2024; 15(7):846. https://doi.org/10.3390/genes15070846

Chicago/Turabian Style

Vera, Brenda, Elly A. Navajas, Pablo Peraza, Beatriz Carracelas, Elize Van Lier, and Gabriel Ciappesoni. 2024. "Genomic Regions Associated with Resistance to Gastrointestinal Parasites in Australian Merino Sheep" Genes 15, no. 7: 846. https://doi.org/10.3390/genes15070846

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop